首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
To model the structural and functional parts of the water oxidizing complex in Photosystem II, a dimeric manganese(II,II) complex (1) was linked to a ruthenium(II)tris-bipyridine (Ru(II)(bpy)(3)) complex via a substituted L-tyrosine, to form the trinuclear complex 2 [J. Inorg. Biochem. 78 (2000) 15]. Flash photolysis of 1 and Ru(II)(bpy)(3) in aqueous solution, in the presence of an electron acceptor, resulted in the stepwise extraction of three electrons by Ru(III)(bpy)(3) from the Mn(2)(II,II) dimer, which then attained the Mn(2)(III,IV) oxidation state. In a similar experiment with compound 2, the dinuclear Mn complex reduced the photo-oxidized Ru moiety via intramolecular electron transfer on each photochemical event. From EPR it was seen that 2 also reached the Mn(2)(III,IV) state. Our data indicate that oxidation from the Mn(2)(II,II) state proceeds stepwise via intermediate formation of Mn(2)(II,III) and Mn(2)(III,III). In the presence of water, cyclic voltammetry showed an additional anodic peak beyond Mn(2)(II,III/III,III) oxidation which was significantly lower than in neat acetonitrile. Assuming that this peak is due to oxidation to Mn(2)(III,IV), this suggests that water is essential for the formation of the Mn(2)(III,IV) oxidation state. Compound 2 is a structural mimic of the water oxidizing complex, in that it links a Mn complex via a tyrosine to a highly oxidizing photosensitizer. Complex 2 also mimics mechanistic aspects of Photosystem II, in that the electron transfer to the photosensitizer is fast and results in several electron extractions from the Mn moiety.  相似文献   

2.
In this short communication we describe the synthesis and the optical and magnetic properties of optically active three dimensional (3D) bimetallic [Cr-Mn] networks [[Delta Cr(III) Delta Mn(II)(ox)(3)][Delta Ru(II)(bpy)(3)]ClO(4)](n)1 - Delta, [[Lambda Cr(III)Lambda Mn(II)(ox)(3)][Lambda Ru(II) (bpy)(3)]ClO(4)](n) 1 - Lambda and [[Delta Cr(III)Delta Mn(II)(ox)(3)][Delta Ru(II)(bpy)(2)p p y]](n) 2 - Delta,[[Lambda Cr(III)Lambda Mn(II)(ox)(3)][Lambda Ru(II)(bpy)(2)ppy]](n) 2 - Lambda (ox = oxalate, bpy = bipyridine, ppy = phenyl-pyridine).  相似文献   

3.
Cyclometalated ruthenium(II) complexes, [Ru(II)(C~N)(N~N)(2)]PF(6) [HC~N=2-phenylpyridine (Hphpy) or 2-(4'-tolyl)pyridine; N~N=2,2'-bipyridine, 1,10-phenanthroline, or 4,4'-dimethyl-2,2'-bipyridine], are rapidly oxidized by H(2)O(2) catalyzed by plant peroxidases to the corresponding Ru(III) species. The commercial isoenzyme C of horseradish peroxidase (HRP-C) and two recently purified peroxidases from sweet potato (SPP) and royal palm tree (RPTP) have been used. The most favorable conditions for the oxidation have been evaluated by varying the pH, buffer, and H(2)O(2) concentrations and the apparent second-order rate constants ( k(app)) have been measured. All the complexes studied are oxidized by HRP-C at similar rates and the rate constants k(app) are identical to those known for the best substrates of HRP-C (10(6)-10(7) M(-1) s(-1)). Both cationic (HRP-C) and anionic (SPP and RPTP) peroxidases show similar catalytic efficiency in the oxidation of the Ru(II) complexes. The mediating capacity of the complexes has been evaluated using the SPP-catalyzed co-oxidation of [Ru(II)(phpy)(bpy)(2)]PF(6) and catechol as a poor peroxidase substrate as an example. The rate of enzyme-catalyzed oxidation of catechol increases more than 10000-fold in the presence of the ruthenium complex. A simple routine for calculating the rate constant k(c) for the oxidation of catechol by the Ru(III) complex generated enzymatically from [Ru(II)(phpy)(bpy)(2)](+) is proposed. It is based on the accepted mechanism of peroxidase catalysis and involves spectrophotometric measurements of the limiting Ru(II) concentration at different concentrations of catechol. The calculated k(c) value of 0.75 M(-1) s(-1) shows that the cyclometalated Ru(II) complexes are efficient mediators in peroxidase catalysis.  相似文献   

4.
The processes that are photoinduced by [Ru(bpz)(3)](2+) (bpz = 2,2'-bipyrazyl) in the presence of Cu/Zn superoxide dismutase (Cu/Zn SOD) are investigated by laser flash photolysis and electron paramagnetic resonance (EPR) spectroscopy; they are compared to those of the system [Ru(bpy)(3)(2+)-Cu/Zn SOD]. Although the mechanism is complicated, primary and secondary reactions can be evidenced. First, the excited [Ru(bpz)(3)](2+) complex is quenched reductively by Cu/Zn SOD with the production of a reduced complex and an oxidized enzyme. The oxidation site of Cu/Zn SOD is proposed to correspond to amino acids located on the surface of the protein. Afterward and only when this reductive electron transfer to the excited complex has produced enough oxidized protein, another electron-transfer process can be evidenced. In this case, however, the charge-transfer process takes place in the other direction, i.e., from the excited complex to the Cu(II) center of the SOD with the formation of Ru(III) and Cu(I) species. This proposed mechanism is supported by the fact that [Ru(bpy)(3)](2+), which is less photo-oxidizing than [Ru(bpz)(3)](2+), exhibits no photoreaction with Cu/Zn SOD. Because Ru(III) species are generated as intermediates with [Ru(bpz)(3)](2+), they are proposed to be responsible for the enhancement of [poly(dG-dC)](2) and [poly(dA-dT)](2) oxidation observed when Cu/Zn SOD is added to the [Ru(bpz)(3)](2+)-DNA system.  相似文献   

5.
A spectroscopic and spectroelectrochemical comparison is made among homo- and heterobimetallic complexes of the form [(bpy)2Ru(BL)Os(byp)2]4+, [(bpy)2Ru(BL)Ru(bpy)2]4+ and [(bpy)2Os(BL)Os(bpy)2]4+ (BL = 2,3,-bis(2′-pyridyl)pyrazzine(dpp),2,3-bis(2′-pyridyl)quinoxaline(dpq) or 2,3-bis(2′-pyridyl)benzoquinoxaline(dpb); bpy = 2,2′-bipyridine). It has been postulated that the spectroscopy of the mixed-metal bimetallic complexes bridged by polyazine bridging ligands can be assigned by comparison to those of the homobimetallic analogs. We have in hand a unique series of complexes where such a postulate can be tested. Utilizing the visible spectra of the homobimetallic Os,Os and Ru,Ru systems, we have been able to generate the spectra of the mixed-metal complexes. Some differences have been seen, particularly in the energy of the Os → dpp 3MLCT. Oxidative spectroelectrochemistry studies on the homobimetallic ruthenium or osmium based systems indicate that upon complete oxidation of both metal centers, transitions in the visible are lost. Hence, partial oxidation of the ruthenium based homobimetallics and Os, Ru mixed-metal bimetallics allows for the direct comparison of the spectroscopic character of the one remaining ruthenium chromophore within these mixed-valence systems. Oxidation to form the Os(III)/Ru(II) species and the Ru(III)/Ru(II) species resulted in similar spectra. This establishes further that the visible spectroscopy of mixed-metal systems of this nature can be accurately interpreted by comparison to the homobimetallic analogs.  相似文献   

6.
A number of ruthenium complexes were tested for their ability to induce filamentation in Escherichia coli. These included monomeric and dimeric complexes with ruthenium in the II or III oxidation states, as well as mixed-valence complexes with ruthenium in the (II,III) oxidation states. In general, dimeric mixed-valence Ru(II,III) complexes were the most active class of compound, although some complexes of this type were relatively inactive. These were pyrazine- or bipyridyl-bridged complexes which are known to involve strong metal-ligand interaction, which stabilizes the Ru(II) oxidation state. Some Ru(III) complexes were also significantly active in induction of filamentous growth in E. coli. One of these was [Ru(NH3)5Cl]Cl2, which did not inhibit electron transport, Mg2+-ATPase activity or DNA synthesis in E. coli, but like [Ru2(NH3)6Br3]Br2 X H2O was a potent inhibitor of respiration-driven calcium transport in the organism. Filament-inducing activity of the complex was reduced in the presence of NaCl, but not in the presence of added Ca2+, ethanol, calcium pantothenate, or E. coli 'division promoting extract'. This behaviour is also similar to that of [Ru2(NH3)6Br3]Br2 X H2O. It is suggested that both complexes may induce filamentation in E. coli by a common mechanism, which may involve interference with calcium metabolism, or a wall or membrane target, rather than interaction with DNA.  相似文献   

7.
《Luminescence》2002,17(2):117-122
The electrogenerated chemiluminescence of Ru(bpy)32+/C2O42? system on a pre‐polarized Au electrode was studied using a potential‐resolved electrochemiluminescence (PRECL) method. Two anodic ECL peaks were observed at 1.22 V (vs. SCE) (EP1), 1.41 V (vs. SCE) (EP2), respectively. The effects of the concentration of oxalate and Ru(bpy)32+, adsorbed sulphur, CO2, O2, pH of the solution and pretreatment of the Au electrode on the two PRECL peaks were examined. The surface state of the pre‐oxidized gold electrode was also studied using the X‐ray photoelectron spectroscopy (XPS) technique. Moreover, comparative studies on i–E and I–E curves were carried out and a possible mechanism involving both the catalytic and the direct electro‐oxidation pathways was proposed for the ECL of Ru(bpy)32+/C2O42? system. EP1 is attributed to the Ru(bpy)32/3+ reaction catalysed by C2O42? to generate Ru(bpy)32+*. EP2 is likely because C2O42? was oxidized at the electrode to form CO2, followed by reaction with Ru(bpy)33+ to generate Ru(bpy)32+*. Copyright © 2002 John Wiley & Sons, Ltd.  相似文献   

8.
Electrochemiluminescence (ECL) based on allantoin and tris(2,2'-bipyridine)ruthenium (II) [Ru(bpy)3 (2+)] was studied in aqueous alkaline buffer solution (pH 11.0). In a flowing system, the eluted allantoin was mixed with 1.0 mmol/L Ru(bpy)3 (2+). When the solution passed through a thin layer flow electrolytic cell equipped with a glassy carbon disc electrode (22.1 mm2), both hydroxyl groups of allantoin and Ru(bpy)3 (2+) were oxidized at the potential of +1.50 V (vs. Ag/AgCl). The luminescence with lambdamax 610 nm caused by the reaction of electrolytically formed Ru(bpy)3 (2+) with alkoxide radical to generate the excited state of Ru(bpy)3 (2+*). A possible ECL process of allantoin in Ru(bpy)3 (2+) alkaline solution has been discussed. In addition, the factors affecting the ECL response of allantoin are also investigated.  相似文献   

9.
The pH-dependent heterometallic complex formation with p-sulfonatothiacalix[4]arene (TCAS) as bridging ligand in aqueous solutions was revealed by the use of spectrophotometry, nuclear magnetic relaxation and fluorimetry methods. The novelty of the structural motif presented is that the appendance of emission metal center ([Ru(bpy)3]2+) is achieved through the cooperative non-covalent interactions with the upper rim of TCAS. The second metal block (Fe(III), Fe(II) and Mn(II)), bound with the lower rim of TCAS in the inner sphere coordination mode is serving as quencher of [Ru(bpy)3]2+ emission. The difference between the complex ability of Fe(III) and Fe(II) ions provides pH conditions for redox-dependent emission of [Ru(bpy)3]2+.  相似文献   

10.
A novel bridging ligand bdptb(2,2'-bis(5,6-diphenyl-1,2,4-triazin-3-yl)-4,4'-bipyridine) and it's chiral diruthenium(II) complex DeltaDelta- and LambdaLambda-[Ru(bpy)2(bdptb)Ru(bpy)2]4+ (Ru2) have been synthesized and characterized by electrospray mass spectra, 1H NMR, UV/visible and circular dichroism spectra. Binding behavior of these dimeric complexes with calf thymus DNA have been investigated by absorption spectra, viscosity measurements, equilibrium dialysis experiments. The electronic absorption spectra hypochromism at the metal-ligand charge transfer of the DeltaDelta- and LambdaLambda-enantiomer are 26.4%, and 40%, and bathochromism of 13.5, and 14 nm in sequence. Equilibrium dialysis experiments results show also the binding-DNA of LambdaLambda-enantiomer is stronger than DeltaDelta-enantiomer. The increasing amounts of the novel dimeric ruthenium(II) complexes on the relative viscosities of calf thymus DNA is smaller than that of the classic intercalators such as [Ru(bpy)2(dppz)]2+ and larger than that of the non-classic intercalators such as Delta-[Ru(phen)3]2+. The experiments suggest the dimeric ruthenium(II) complex may be bound to DNA by groove binder.  相似文献   

11.
Metal-to-ligand charge transfer photolysis of the ruthenium(II) pyrazine complex Ru(NH3)5pz2+ (I) in pH 7.4 oxygenated phosphate buffer solution generates the Ru(III) analog Ru(NH3)5pz3+ plus the reactive oxygen species singlet oxygen and superoxide. Based on the very short MLCT lifetime (re-measured as approximately 250 ps in D2O) of I* and the quantum yield for singlet oxygen formation (0.01 for aerated D2O) the rate constant for oxygen quenching of I* was calculated to be approximately (3+/-1)x10(10) M-1 s-1.  相似文献   

12.
The new complex, [RuII(bpy)2(4-HCOO-4′-pyCH2 NHCO-bpy)](PF6)2 · 3H2O (1), where 4-HCOO-4′-pyCH2NHCO-bpy is 4-(carboxylic acid)-4′-pyrid-2-ylmethylamido-2,2′-bipyridine, has been synthesised from [Ru(bpy)2(H2dcbpy)](PF6)2 (H2dcbpy is 4,4′-(dicarboxylic acid)-2,2′-bipyridine) and characterised by elemental analysis and spectroscopic methods. An X-ray crystal structure determination of the trihydrate of the [Ru(bpy)2(H2dcbpy)](PF6)2 precursor is reported, since it represented a different solvate to an existing structure. The structure shows a distorted octahedral arrangement of the ligands around the ruthenium(II) centre and is consistent with the carboxyl groups being protonated. A comparative study of the electrochemical and photophysical properties of [RuII(bpy)2(4-HCOO-4′-pyCH2NHCO-bpy)]2+ (1), [Ru(bpy)2(H2dcbpy)]2+ (2), [Ru(bpy)3]2+ (3), [Ru(bpy)2Cl2] (4) and [Ru(bpy)2Cl2]+ (5) was then undertaken to determine their variation upon changing the ligands occupying two of the six ruthenium(II) coordination sites. The ruthenium(II) complexes exhibit intense ligand centred (LC) transition bands in the UV region, and broad MLCT bands in the visible region. The ruthenium(III) complex, 5, displayed overlapping LC bands in the UV region and a LMCT band in the visible. 1, 2 and 3 were found, via cyclic voltammetry at a glassy carbon electrode, to exhibit very positive reversible formal potentials of 996, 992 and 893 mV (versus Fc/Fc+) respectively for the Ru(III)/Ru(II) half-cell reaction. As expected the reversible potential derived from oxidation of 4 (−77 mV (versus Fc/Fc+)) was in excellent agreement with that found via reduction of 5 (−84 mV (versus Fc/Fc+)). Spectroelectrochemical experiments in an optically transparent thin-layer electrochemical cell configuration allowed UV-Vis spectra of the Ru(III) redox state to be obtained for 1, 2, 3 and 4 and also confirmed that 5 was the product of oxidative bulk electrolysis of 4. These spectrochemical measurements also confirmed that the oxidation of all Ru(II) complexes and reduction of the corresponding Ru(III) complex are fully reversible in both the chemical and electrochemical senses.  相似文献   

13.
A novel preparation method for surfactant-MnP-Mn(II) ternary complex utilizing water-in-oil emulsions has been developed. The surfactant-MnP complex was spectroscopically characterized, strongly suggesting that the heme environment of the surfactant-MnP complex in benzene is identical to that of native MnP in the aqueous buffer. o-Phenylenediamine oxidation catalyzed by the surfactant-MnP-Mn(II) ternary complex was performed in benzene. The ternary complex efficiently catalyzed the oxidation, and the complex was catalytically stable. Kinetic experiments revealed that the reaction mechanism was as follows: MnP is oxidized by H(2)O(2) and the oxidized intermediate catalyzes the oxidation of Mn(II) to Mn(III) and the latter, after complexed with malonate, readily oxidizes o-PDA inside the complex. Thus, the organic substrate o-PDA, but not Mn(III), shuttled between the surfactant-MnP-Mn(II) ternary complex and organic solvent.  相似文献   

14.
The characterization of a polyclonal antibody produced via immunization with an [Os(bpy)(2)dcbpy] hapten is described. Bpy is 2,2'-bipyridine and dcbpy is 2,2'-bipyridine-4,4'-dicarboxylic acid. The cross-reactivity of the antibody for the Ru(II) analogue of the hapten was also investigated. Large increases in the emission and luminescent lifetime of a series of Os and Ru complexes were observed on binding of the antibody. Association equilibrium constants were derived from luminescence titration data and were found to be 5.6 x 10(8) and 5.0 x 10(8)M(-1) for [Os(bpy)(2)dcbpy] and [Ru(bpy)(2)dcbpy], respectively. Spectroscopic changes were likely due to the exclusion of H(2)O from the complex/antibody binding cleft and blocking of vibrational relaxation pathways of the Os/Ru excited state. D(2)O/H(2)O experiments confirmed that the antibody protected approx. 82% of [Os(bpy)(2)dcbpy] and 80% of [Ru(bpy)(2)dcbpy] from excited state deactivation by the aqueous solvent.  相似文献   

15.
A comparative study of the binding of square planar cis- and trans-[Pt(NH3)2Cl2] complexes and the octahedral [Ru(NH3)5(H2O)]3+ complex to tRNAphe from yeast was carried out by X-ray crystallography. Both of the carcinostatic compounds, cis-[Pt(NH3)2Cl2] and [Ru(NH3)5(H2O)]3+ show similarities in their mode of binding to tRNA. These complexes bind specifically to the N(7) positions of guanines G15 and G18 in the dihydrouridine loop. [Ru(NH3)5(H2O)]3+ has an additional binding site at N(7) of residue G1 after extensive soaking times (58 days). A noncovalent binding site for ruthenium is also observed in the deep groove of the acceptor stem helix with shorter (25 days) soaking time. The major binding site for the inactive trans-[Pt(NH3)Cl2] complex is at the N(1) position of residue A73, with minor trans-Pt binding sites at the N(7) positions of residues Gm34, G18 and G43. The similarities in the binding modes of cis-[Pt(NH3)2Cl2] and [Ru(NH3)5(H2O)]3+ are expected to be related to their carcinostatic properties.  相似文献   

16.
Ruthenium(II) tris(bipyridine)-centered poly(ethylenimine) (Ru PEI) was synthesized via acid hydrolysis of Ru tris(bipyridine)-centered poly(2-ethyl-2-oxazoline) (Ru PEOX), and the luminescence, DNA entrapment, and transfection efficiencies were evaluated. Emission maxima for Ru PEI samples are red-shifted compared to Ru PEOX precursors, and the luminescence lifetimes are shorter in both methanol and aqueous solutions. Slower oxygen quenching of Ru PEOX and Ru PEI luminescence versus [Ru(bpy)3]Cl2 (bpy = bipyridine) is attributed to polymer shielding effects. Ru PEI luminescence is similar in the presence and absence of DNA. Ru PEI (7900 Da) and linear PEI (L-PEI; 22,000 Da) fully entrapped DNA (5.4 kb; pcDNA) at an N/P ratio of 2. LNCaP prostate cancer cells were transfected with a plasmid encoding for green fluorescent protein using Ru PEI and L-PEI vectors for comparison. For N/P = 48, the transfection efficiency for Ru PEI was approximately 50% relative to that of L-PEI.  相似文献   

17.
A new Ru(II) complex of [Ru(bpy)(2)(Hpip)](2+) {bpy = 2,2'bipyridine; Hppip = 2-(4-(pyridin-2-yl)phenyl)-1H-imidazo[4,5-f][1,10]phenanthroline} has been synthesized by grafting of 2-pyridyl to parent complex [Ru(bpy)(2)(Hpip)](2+) {Hppip = 2-(4-phenyl)-1H-imidazo[4,5-f][1,10]phenanthroline}. The acid-base properties of [Ru(bpy)(2)(Hpip)](2+) studied by UV-visible and luminescence spectrophotometric pH titrations, revealed off-on-off luminescence switching of [Ru(bpy)(2)(Hpip)](2+) that was driven by the protonation/deprotonation of the imidazolyl and the pyridyl moieties. The complex was demonstrated to be a DNA intercalator with an intrinsic DNA binding constant of (5.56 ± 0.2) x 10(5) M-1 in buffered 50 mM NaCl, as evidenced by UV-visible and luminescence titrations, reverse salt effect, DNA competitive binding with ethidium bromide, steady-state emission quenching by [Fe(CN)6]4-, DNA melting experiments and viscosity measurements. The density functional theory method was also used to calculate geometric/electronic structures of the complex in an effort to understand the DNA binding properties. All the studies indicated that the introduction of 2-pyridyl onto Hpip ligand is more favorable for extension of conjugate plane of the main ligand than that of phenyl, and for greatly enhanced ct-DNA binding affinity accordingly.  相似文献   

18.
A new Ru(II) complex, [Ru(bpy)(2)(dhipH3)](ClO4)(2) (in which bpy=2,2'-bipyridine, dhipH(3)=3,4-dihydroxy-imidado[4,5-f][1,10]-phenanthroline), was synthesized and characterized, and the pH effect on the emission spectra of the complex was studied. The interaction of the complex with calf thymus DNA was investigated by UV-visible and emission spectroscopy, and viscosity measurements. The results suggest that the complex acted as a sensitive luminescent pH sensor and a strong ct-DNA intercalator with an intrinsic binding constant of (4.0+/-0.7) x 10(5) M(-1) in buffered 50 mM NaCl.  相似文献   

19.
The manganese peroxidase (MnP), from the lignin-degrading fungus Phanerochaete chrysosporium, an H2O2-dependent heme enzyme, oxidizes a variety of organic compounds but only in the presence of Mn(II). The homogeneous enzyme rapidly oxidizes Mn(II) to Mn(III) with a pH optimum of 5.0; the latter was detected by the characteristic spectrum of its lactate complex. In the presence of H2O2 the enzyme oxidizes Mn(II) significantly faster than it oxidizes all other substrates. Addition of 1 M equivalent of H2O2 to the native enzyme in 20 mM Na-succinate, pH 4.5, yields MnP compound II, characterized by a Soret maximum at 416 nm. Subsequent addition of 1 M equivalent of Mn(II) to the compound II form of the enzyme results in its rapid reduction to the native Fe3+ species. Mn(III)-lactate oxidizes all of the compounds which are oxidized by the enzymatic system. The relative rates of oxidation of various substrates by the enzymatic and chemical systems are similar. In addition, when separated from the polymeric dye Poly B by a semipermeable membrane, the enzyme in the presence of Mn(II)-lactate and H2O2 oxidizes the substrate. All of these results indicate that the enzyme oxidizes Mn(II) to Mn(III) and that the Mn(III) complexed to lactate or other alpha-hydroxy acids acts as an obligatory oxidation intermediate in the oxidation of various dyes and lignin model compounds. In the absence of exogenous H2O2, the Mn-peroxidase oxidized NADH to NAD+, generating H2O2 in the process. The H2O2 generated by the oxidation of NADH could be utilized by the enzyme to oxidize a variety of other substrates.  相似文献   

20.
Ascomycetes that can deposit Mn(III, IV) oxides are widespread in aquatic and soil environments, yet the mechanism(s) involved in Mn oxide deposition remains unclear. A Mn(II)-oxidizing ascomycete, Acremonium sp. strain KR21-2, produced a Mn oxide phase with filamentous nanostructures. X-ray absorption near-edge structure (XANES) spectroscopy showed that the Mn phase was primarily Mn(IV). We purified to homogeneity a laccase-like enzyme with Mn(II) oxidase activity from cultures of strain KR21-2. The purified enzyme oxidized Mn(II) to yield suspended Mn particles; XANES spectra indicated that Mn(II) had been converted to Mn(IV). The pH optimum for Mn(II) oxidation was 7.0, and the apparent half-saturation constant was 0.20 mM. The enzyme oxidized ABTS [2,2'-azinobis(3-ethylbenzothiazoline-6-sulfonic acid)] (pH optimum, 5.5; Km, 1.2 mM) and contained two copper atoms per molecule. Moreover, the N-terminal amino acid sequence (residues 3 to 25) was 61% identical with the corresponding sequence of an Acremonium polyphenol oxidase and 57% identical with that of a Myrothecium bilirubin oxidase. These results provide the first evidence that a fungal multicopper oxidase can convert Mn(II) to Mn(IV) oxide. The present study reinforces the notion of the contribution of multicopper oxidase to microbially mediated precipitation of Mn oxides and suggests that Acremonium sp. strain KR21-2 is a good model for understanding the oxidation of Mn in diverse ascomycetes.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号